• Aucun résultat trouvé

Multiscale Finite Element approach for “weakly” random problems and related issues

N/A
N/A
Protected

Academic year: 2022

Partager "Multiscale Finite Element approach for “weakly” random problems and related issues"

Copied!
44
0
0

Texte intégral

(1)

DOI:10.1051/m2an/2013122 www.esaim-m2an.org

MULTISCALE FINITE ELEMENT APPROACH

FOR “WEAKLY” RANDOM PROBLEMS AND RELATED ISSUES

Claude Le Bris

1,2

, Fr´ ed´ eric Legoll

2,3

and Florian Thomines

2,3

Abstract.We address multiscale elliptic problems with random coefficients that are a perturbation of multiscale deterministic problems. Our approach consists in taking benefit of the perturbative context to suitably modify the classical Finite Element basis into a deterministic multiscale Finite Element basis. The latter essentially shares the same approximation properties as a multiscale Finite Element basis directly generated on the random problem. The specific reference method that we use is the Multiscale Finite Element Method. Using numerical experiments, we demonstrate the efficiency of our approach and the computational speed-up with respect to a more standard approach. In the stationary setting, we provide a complete analysis of the approach, extending that available for the deterministic periodic setting.

Mathematics Subject Classification. 35B27, 65M60, 65M12, 35R60, 60H.

Received November 7, 2011.

Published online April 8, 2014.

1. Overview of our approach and results

The Multiscale Finite Element Method (henceforth abbreviated as MsFEM) is a popular numerical approach for multiscale problems (see [3,11,18,20,30–32,37–39]). It consists in a Galerkin approximation of the original problem over a finite dimensional space generated by basis functions that are specificallyadaptedto the problem under consideration.

This approach is popular for a twofold reason. First, its use is not restricted to multiscale problems that converge to a homogenized problem in the limit of vanishing ratio between the small scale and the macroscopic scale. It may be applied to much more general situations. Second, when the problem does converge to a homog- enization problem, the MsFEM approach is meant to approximate the solution of the problem with the small scaleεat its actual small value and not “only” in the asymptotic regimeε→0, which is the regime addressed by homogenization theory.

Keywords and phrases.Weakly stochastic homogenization, finite elements, Galerkin methods, highly oscillatory PDE.

1 CERMICS, ´Ecole Nationale des Ponts et Chauss´ees, Universit´e Paris-Est, 6 et 8 avenue Blaise Pascal, 77455 Marne-La-Vall´ee Cedex 2, France.lebris@cermics.enpc.fr

2 INRIA Rocquencourt, MICMAC team-project, Domaine de Voluceau, B.P. 105, 78153 Le Chesnay Cedex, France.

3 Laboratoire Navier, ´Ecole Nationale des Ponts et Chauss´ees, Universit´e Paris-Est, 6 et 8 avenue Blaise Pascal, 77455 Marne- La-Vall´ee Cedex 2, France.legoll@lami.enpc.fr;thominef@lami.enpc.fr

Article published by EDP Sciences c EDP Sciences, SMAI 2014

(2)

ET AL.

To fix the ideas, consider the problem of finding uεsolving

−div [Aε∇uε] =f inD, uε= 0 on∂D, (1.1) on a bounded domain D ⊂Rd, withf ∈L2(D), and whereAε is a uniformly bounded, coercive matrix that varies at scaleε. A standard Finite Element Method (FEM) would require a space discretization of the domain at the scaleεin order to capture the oscillations ofuεat scaleε. This is prohibitively expensive. The MsFEM aims at accurately approximating uε using a limited number of degrees of freedom. It does not require the matrixAεto be periodic (namelyAε(x) =Aper(x/ε) for a fixed periodic matrix Aper) or stationary.

We now briefly describe the approach and present the aim of this article. Starting from a coarse mesh Th

with a standard (sayP1) Finite Element basis set of functions φ0iL

i=1, generating the associated space Vh:= span(φ0i, i= 1, . . . , L),

we first numerically build the MsFEM basis functionsφεi. Several definitions of these basis functions have been proposed in the literature (yielding different numerical methods), and we detail this in the sequel (seee.g.(2.7)- (2.8)-(2.9)). For the moment, it is sufficient to know that, to eachφ0i, which varies at the macroscopic scale, is associated a function φεi, with variations at the scale ε. In practice, φεi is numerically computed (in fact, pre-computed), using the specificities of the problem addressed. These highly oscillatory functions φεi generate the finite dimensional space

Wh:= span(φεi, i= 1, . . . , L).

Note thatWh andVh share the same dimension.

We next define the MsFEM solution uM using a Galerkin approximation of (1.1) on Wh, instead of Vh. Again, details will be given below. The MsFEM solutionuM provided by the approach reads

uM(x) = L i=1

(UM)i φεi(x),

for some coefficients{(UM)i}Li=1. Of course, these coefficients depend onε, but this dependency is kept implicit in the sequel.

We now turn our attention to the stochastic problem

−div [Aε(·, ω)∇uε(·, ω)] =f inD, uε(·, ω) = 0 on∂D, (1.2) and a typical quantity of interestE[uε(x,·)], which is traditionally approximated using a Monte Carlo method.

Introducing a set of M realizations of the stochastic matrix {Aε,m}1≤m≤M, a direct, na¨ıve application of the MsFEM paradigm would consist in first computing for each realization m the stochastic MsFEM basis functionsφε,mi (x, ω), next performing a Galerkin approximation of (1.2) using this MsFEM basis set to compute {umM(x, ω)}1≤m≤M, and eventually approximatingE[uε(x,·)] by

E[uε(x,·)]≈ 1 M

M m=1

umM(x, ω).

Such an approach is unpractical because of the prohibitively expensive computational load.

To reduce the computational cost and make the MsFEM approach practical in such a stochastic context, a natural idea we investigate in this article is to consider a less generic setting, for which a dedicated, more computationally affordable approach, can be designed. One possibility is to consider matricesAε(x, ω)≡Aε(x)+

B(x, ω) in (1.2) that are not highly oscillatory in their stochastic part. In such cases, dedicated approaches have been proposed, we refer to [35] for more details. Another approach is to reduce the number of Monte−Carlo

(3)

simulations used for the computation of the multiscale basis functions. In [1,26], the authors assume that their coefficient can be written as a Karhunen-Lo`eve type expansion, and apply a collocation method to a priori choose some sparse realizations for which they compute the multiscale basis functions.

In this article, we consider one of the many alternate variants of problem (1.2). We suppose that Aε(x, ω) is highly oscillatory in both its deterministic and stochastic components, but that it is a perturbation of a deterministic matrix. More precisely, we assume that

Aε(x, ω)≡Aεη(x, ω) =Aε0(x) +ηAε1(x, ω), (1.3) whereAε0 is a deterministic matrix andηis a small deterministic parameter. This model may be well suited for heterogeneous materials (or, more generally, media) that, although not periodic, arenot fullystochastic, in the sense that they may be considered as aperturbationof a deterministic material. We call this setting theweakly stochastic setting. Note that many practical situations, involving actual materials or media, can be considered, at a good level of approximation, as perturbations of a deterministic (often periodic) setting (seee.g.[41]).

In a series of recent works (see [14,15,25] and [6–8]; see also [5] for a unified presentation), we have considered such a setting, in the context of homogenization theory (the matrixAεη(x, ω) in (1.2)−(1.3) readsAεη(x, ω) = Aη(x/ε, ω) for a stationary matrixAη(x, ω), which is, in a sense to be made precise, a perturbation of a periodic matrix). We have shown there that, in such a case, the workload for computing the homogenized solution is significantly lighter than for generic stochastic homogenization, and actually comparable to the workload for periodic homogenization. We will show in the sequel that the MsFEM can be adapted to this weakly stochastic setting, providing an approximation of the solutionuεη to (1.2)−(1.3), for fixedε, at a much smaller computational cost than the direct approach.

The main idea of our proposed approach is to compute a set of deterministicMsFEM basis functions using Aε0, the deterministic part of Aεη in the expansion (1.3), and then to perform Monte Carlo realizations at the macroscale level using a set ofMrealizations of the random matrix

Aε,mη (x, ω)

1≤m≤M (see Sect.2 for a detailed presentation). Note that, for each of these realizations, we solve theoriginalproblem, with thecomplete matrixAεη, and not only its deterministic part. Only the basis set is taken deterministic. By construction, the approach provides an approximation

uS(x, ω) = L i=1

(US(ω))i φεi(x)

ofuεη(x, ω), where the basis functionsφεi aredeterministic. These basis functions are computed only once, hence the cost to compute {umS(x, ω)}1≤m≤M is much smaller than the cost to compute {umM(x, ω)}1≤m≤M. This is especially true if (1.2) has to be solved for many right-hand sidesf. We expect that this approximationuS is as accurate asuM for smallη. We show below that this is indeed the case, whenAεη is a perturbation ofAε0(see Sect.3for numerical tests).

We would like to note that the MsFEM is not the only multiscale technique based on finite elements. The bottom line of our approach, consisting of generating suitable multiscale functions for the discretization of a weakly stochastic problem, using for this purpose the deterministic reference problem, can in principle be applied to other multiscale techniques. Another popular technique is the HMM method [27–29], for which our approach could in principle be easily adapted.

In the numerical tests reported on in Section3, we compare, in theH1 norm,uεη (the exact solution to (1.2) with the matrixAε≡Aεη given by (1.3)) withuS (the solution provided by our approach) anduM (the solution provided by the ideal, expensive approach). The quantityuεη−uMH1(D)represents the best possible accuracy that we can achieve, in the sense that our approach inherits the limitations of the MsFEM approach. We thus cannot expect our approximation uS to be more accurate than uM. We can only hope to compute an approximation of comparable quality with a much reduced workload. The numerical results we obtain confirm that, for small η in (1.3), the quantityuS−uεηH1(D) is of the same order of magnitude asuM−uεηH1(D),

(4)

ET AL.

although, we repeat it, the computational cost to compute uS is much smaller than that to computeuM. In Section3, we also show the advantage of performing Monte Carlo realizations at the macroscale level (using the random matrixAε,mη ) over solving a deterministic macroscale problem with only the deterministic part ofAεη.

We next derive error bounds for our approach in Section 4. We recall that, in the deterministic setting, a classical context for proving convergence of the MsFEM approach is the case when, in the reference problem (1.1), the matrix reads Aε(x) = Aper

x ε

for a fixed periodic matrix Aper. Likewise, to be able to perform our theoretical analysis in the stochastic setting, we assume in Section 4 that Aεη(x, ω) = Aη

x ε, ω

for a fixed stationary random matrixAη. The problem (1.2)(1.3) then admits a homogenized limit whenεvanishes.

Our proof follows the same lines as that in the deterministic setting, which we now briefly review (see the introduction of Sect.4for more details on the structure of the proof). The MsFEM is a Galerkin approximation, that we assume momentarily, for the sake of clarity, to be aconformingapproximation (this is indeed the case when, for defining the highly oscillatory basis functionsφεi, oversampling isnotused). The error is then estimated using the C´ea lemma:

uε−uMH1(D)≤C inf

vh∈Wh

uε−vhH1(D),

where uε is the solution to the reference deterministic highly oscillatory problem (1.1), uM is the MsFEM solution and C is a constant independent of ε and h. Taking advantage of the homogenization setting, we introduce the two-scale expansion

vε=u+ε d i=1

w0e

i

· ε

iu

ofuε, whereuis the homogenized solution,w0ei is the periodic corrector associated toeiRd, andiudenotes the partial derivative ∂u

∂xi. We next write uε−uMH1(D)≤C

uε−vεH1(D)+ inf

vh∈Wh

vε−vhH1(D)

.

The first term in the above right-hand side is estimated using standard homogenization results on therate of convergence of vε−uε. To estimate the second term, one considers a suitably chosen element vh ∈ Wh, for whichvε−vhH1(D)can be directly bounded.

Following the same strategy in our stochastic setting, we estimate the distance between the solutionuεη to the reference stochastic problem (1.2)(1.3) and the weakly stochastic MsFEM solutionuS as

uεη(·, ω)−uS(·, ω)H1(D)≤C

uεη(·, ω)−vεη(·, ω)H1(D)+ inf

vh∈Wh

vηε(·, ω)−vhH1(D)

.

We observe that a key ingredient for the proof is the rate of convergence of the difference between the reference solution uεη and its two-scale expansion vεη. Such a result is classical in periodic homogenization, but, to the best of our knowledge, open in the general stationary case (in dimensions higher than one). One only knows thatuεη−vηεvanishes (in some appropriate norm) whenε→0. However, in the particular case whenAεηis only weaklystochastic, it is possible to obtain such a result, as we have shown in [42]. Hence, exploiting the specificity of our weakly stochastic setting, we are able to obtain (see our main result, Thm.4.5and estimate (4.37)):

E

uεη−uS2H1 h

≤C

ε+h+ε h+η

ε h

d/2

ln(N(h)) +η+η2C(η)

, where · H1

h is a brokenH1norm,C is a constant independent ofε,handη,Cis a bounded function asηgoes to 0, and N(h) is the number of elements in the mesh (roughly of order h−d in dimensiond). As is often the

(5)

case in the deterministic setting, we use here (both for our numerical tests and in the analysis) the oversampling technique. Consequently, the basis functions φεi do not belong to H01(D), hence the use of a brokenH1 norm in the above estimate. As we point out below, when η = 0 in (1.3), our approach reduces to the standard deterministic MsFEM (with oversampling), and the above estimates then agree with those proved in [32].

This article is organized as follows. First, in Section 2, we describe the MsFEM approach. For consistency, we begin by the deterministic setting in Section 2.1, and point out there that the direct adaptation to the general stochastic setting yields a prohibitively expensive approach. The adaptation of the approach to the weakly stochastic settingis described in Section 2.2. We next turn to numerical simulations, in Section3. Some procedures to efficiently implement the approach are first described in Section 3.1. We next consider a one- dimensional test (see Sect.3.2), which is useful for several reasons. First, it allows to calibrate some numerical parameters, such as the number M of independent realizations when estimating the exact expectation by an empirical mean. Second, we assess the accuracy of our approach with respect to the magnitude of η. We demonstrate there thatη does not have to be extremely small for our method to be very efficient. For instance, on the test case considered in Section 3.2, we show that our approach is as accurate as the expensive, direct approach as soon asη is such that

ηaε1 aε0

L(R×Ω)

is equal to or smaller than 0.1,

whereaε0is the deterministic component of the diffusion coefficientaεη andηaε1 is the stochastic component (see expansion (1.3)). On the other hand, as pointed out above, our approach is not meant to address the regime whenη 1. Lastly, we also assess the accuracy of our approach with respect to the presence of frequencies in the random coefficient aεη that are not taken into account in the MsFEM basis set. We next turn to two test cases in dimension two, where we observe that our approach performs as well as in the one-dimensional case (see Sect. 3.3). In particular, in Section 3.3.2, we successfully address a classical test-case of the literature. In Section3.4, we compare our approach with a fully deterministic approach. All the information about variance is lost when using the latter approach. In contrast, using our approach, we show that we can accurately approximate some quantities of interest which are random in nature, such as the variance of the solution.

Section 4 is devoted to the analysis of the approach, in the homogenization setting (i.e. when the matrix in (1.2) reads Aε(x, ω) = A

x ε, ω

where A is stationary). Our main result, Theorem 4.5, is presented in Section 4.1, and proved in Section 4.2. The proofs of some technical results are collected in Appendix A. In addition, we specifically consider the one dimensional case in Section4.3.

2. MsFEM-type approaches

For consistency and the convenience of the reader, we present in this section the MsFEM approach to solve the original elliptic problem (1.1). For clarity, we begin by presenting the approach in a deterministic setting.

The reader familiar with the MsFEM may easily skip this section and directly proceed to Section2.2, where we present our approach in a weakly stochastic setting.

2.1. Description in a classical deterministic setting

Letuε∈H1(D) be the solution to (1.1), where the matrixAε(L(D))d×d satisfies the standard coercivity condition: there exists two constantsa+≥a >0 such that, for anyε,

∀ξ∈Rd, a|ξ|2≤ξTAε(x)ξa.e. inD and AεL(D)≤a+.

Note that the MsFEM approach is not restricted to the periodic setting. We therefore do not assume that Aε(x) =Aper(x/ε) for a fixed periodic matrix Aper.

(6)

ET AL.

The MsFEM approach consists in performing a variational approximation of (1.1) where the basis functions are precomputed and encode the fast oscillations present in (1.1). In the sequel we argue on the following formulation, equivalent to (1.1):

Finduε∈H01(D) such that, for any v∈H01(D), Aε(uε, v) =b(v), (2.1) where

Aε(u, v) =

D(∇v(x))TAε(x)∇u(x) dx and b(v) =

Df(x)v(x) dx.

The MsFEM is a three-step approach:

1. introduce a standard discretization of the domain D using a coarse mesh as compared to the small scale oscillations ofAε;

2. for each elementKof the coarse mesh, compute the basis functionφε,Ki as the solution of an elliptic equation posed inK(seee.g.(2.7)(2.8)(2.9) below);

3. solve the Galerkin approximation of (2.1), for the set of basis functions defined at Step 2.

The advantage of the approach is that, for the same accuracy of the approximation as that provided by a standard FEM, the macroscale mesh can be chosen sufficiently coarse so that the resulting discretized problem has a limited number of degrees of freedom, and may thus be computationally solved inexpensively. This is observed in practice [37], and proven by a theoretical analysis (see [32,38]) when the problem (2.1) admits a homogenized limit. See also [30] and references therein.

To further illustrate this fact, we reproduce here a simple one-dimensional analysis we borrow from A. Lozinski (see [44], Chap. 6 and [17]). This analysis explains remarkably well the interest of the approach, and, in contrast to [32,38], is not restricted to a homogenization setting. Consider the one-dimensional domainD= (0,1) and the reference problem

Lu=f, u(0) =u(1) = 0,

for the operatorLu:=(νu), wheref ∈L2(0,1) andν∈L(0,1) withν(x)≥νmin>0 almost everywhere on (0,1). The functionν may have oscillations at a small scale. The associated weak formulation reads

Find u∈H01(0,1) such that, for anyv∈H01(0,1), a(u, v) =b(v), (2.2) with

a(u, v) = 1

0

ν(x)u(x)v(x) dx and b(v) = 1

0

f(x)v(x) dx.

We now introduce the nodes 0 = x0 < x1 < · · · < xL = 1 that define the elements Ki = [xi−1, xi]. Let h= max|xi−xi−1|be the mesh size. The multiscale finite element space

Wh=

vh∈C0(0,1) such that Lvh= 0 on eachKi

, (2.3)

defined using the operatorL, isadaptedto the problem under study. We next proceed with a Galerkin approx- imation of (2.2) using the spaceWh:

Find uh∈ Wh such that, for any vh∈ Wh, a(uh, vh) =b(vh).

The solutionuh then satisfies

u−uhE h π√

νminfL2(0,1) (2.4)

where · E =

a(·,·) is the energy norm. The proof of this estimate goes as follows. By definition of uand uh, we havea(u−uh, vh) = 0 for anyvh ∈ Wh. Hence,uhis the orthogonal projection ofuonWh according to the scalar producta(·,·). Since · E is the norm associated to that scalar product, we have

u−uhE= inf

vh∈Wh

u−vhE. (2.5)

(7)

Choosevh to be the finite element interpolant ofu, which is defined by vh(xi) =u(xi) for anyi= 0,1, . . . , L, and consider the interpolation errore=u−vh. On each elementKi, we have, precisely because the spaceWh is defined as (2.3),

Le=−(νe)=f with e(xi−1) =e(xi) = 0.

We multiply bye, integrate by part and obtain xi

xi−1

ν(x)|e(x)|2dx= xi

xi−1

f(x)e(x) dx≤ fL2(Ki)eL2(Ki). (2.6) Sinceevanishes on the boundary ofKi, the Poincar´e inequality with the best constant (xi−xi−1)/πyields

eL2(Ki)≤xi−xi−1

π eL2(Ki) h π√

νmin xi

xi−1

ν(x)|e(x)|2dx 1/2

. By substitution in (2.6), we obtain

xi

xi−1

ν(x)|e(x)|2dx h2

π2νminf2L2(Ki).

Summing over the elements and using (2.5) yields (2.4). Using again thatν is bounded from below, we deduce from (2.4) that

u−uhH1(0,1) h

CDπ νminfL2(0,1),

whereCDis the Poincar´e constant of the domainD= (0,1). As pointed out in Chapter 6 of [44], the interest of the above estimate (or of estimate (2.4)) lies in the fact that the constant in the right-hand side only depends onν throughνmin, and remains the same even ifν oscillates at a small scale. In contrast, for a standard finite element method, the error is also proportional toh, but with a constant that depends on theH2 norm of the exact solutionu. With a standard finite element spaceWh, we indeed classically deduce by C´ea’s lemma that

u−uhH1(0,1) νL(0,1)

CDνmin inf

vh∈Wh

u−vhH1(0,1)= νL(0,1)

CDνmin u−RhuH1(0,1),

where CD is the Poincar´e constant of the domainD= (0,1), andRhuis the projection ofuonWh according to theH1 scalar product. We thus obtain that

u−uhH1(0,1)≤ChνL(0,1)

νmin uL2(0,1),

where C is independent from the functionsν andu. If ν oscillates at a small scale (e.g. ν(x) = ν(x/ε) for a fixed functionν), theH2norm ofumay be large (of the order ofε−1). A FEM approach then requireshto be smaller thanεto reach a good accuracy.

We conclude this illustration by noting that such a general analysis of the MsFEM approach is not available in dimensiond≥2. The analysis presented in [32,38], which is performed without any restriction on the dimension, additionally assumes that the matrixAε in (2.1) readsAε(x) =Aper(x/ε) for a fixedperiodic matrixAper.

We now describe the MsFEM in a multidimensional setting.

2.1.1. Definition of the coarse mesh

For simplicity (see Rem.2.1below), we consider a classicalP1 discretization of the domainD. We denote by Th the corresponding mesh, withLnodes. Let φ0i, i= 1, . . . , L, be the basis functions. We introduce the finite element space

Vh:= span(φ0i, i= 1, . . . , L),

(8)

ET AL.

Figure 1. Definition ofS(in 2D for clarity).

and define the restriction

φ0,Ki :=φ0i

K

of these functions in each elementK.

Remark 2.1. We refer to [3] for a presentation of a MsFEM method that usesP2 macroscale basis functions.

2.1.2. Definition of the MsFEM basis

Several definitions of the MsFEM basis functions have been proposed in the literature (see e.g. [3,30,32,37–39]). They all follow the same pattern but they give rise to various methods. We present in the following the particular method that we have implemented. It makes use of the oversampling technique introduced in [37] and developed in [36].

For any elementK, we consider a domainSK (see Fig.1), obtained from K by an homothetic transfor- mation of center the centroid ofK, and of ratio larger than 1.

LetxSj denote the coordinate of the vertexj of the domainS. For any vertexiofS, we introduce the affine functionχ0,Si (defined onS) that satisfies the conditionχ0,Si (xSj) =δij for allj. Letχε,Si ∈H1(S) be the unique solution to the problem

−div

Aε(x)∇χε,Si (x)

= 0 inS, χε,Si =χ0,Si on∂S, (2.7) which, in practice, is numerically solved e.g. using a finite element method with a mesh size adapted to the small scaleε. We then define the local basis functions

φε,Ki = d+1 j=1

αij χε,Sj

K (2.8)

as linear combinations of the restrictions ofχε,Si onK, withαij chosen such that

∀1≤i, j≤d+ 1, φ0,Ki (xKj ) =

d+1

j=1

αijχ0,Sj (xKj ) =δij, (2.9)

wherexKj denotes the coordinate of thejth vertex of the elementK. Note that the condition (2.9) is enforced on the function φ0,Ki , and not onφε,Ki . The coefficientsαij are consequently independent fromε. Asφ0,Ki and

(9)

χ0,Sj

K are both affine onK, condition (2.9) implies that

∀1≤i≤d+ 1, ∀x∈K, φ0,Ki (x) =

d+1

j=1

αijχ0,Sj (x). (2.10)

We next introduce the functionsφεi defined on Dby φεi|K=φε,Ki for all elementsK.

Note that the problems (2.7), indexed by S, are all independent from one another. They may be solved in parallel.

2.1.3. Macroscale problem

We now introduce the finite dimensional space

Wh:= span(φεi, i= 1, . . . , L), and proceed with the approximation

Find uM ∈ Wh such that, for any v∈ Wh, Ahε(uM, v) =b(v), (2.11) of (2.1), where

Ahε(u, v) =

K

K(∇v(x))TAε(x)∇u(x) dx and b(v) =

Df(x)v(x) dx.

Observe thatφεi has jumps across the edges of the triangulation (due to the use of the oversampling technique), henceWh ⊂H1(D), thus the broken integral used to defineAhε(u, v). On the other hand, sinceWh ⊂L2(D), the linear formbis well defined forv∈ Wh. The formulation (2.11) is anon-conformingGalerkin approximation of (2.1). This brings additional error terms in the error estimation (see Lem.4.7in Section4). On another note, remark that the dimension of Wh is equal to L. The formulation (2.11) hence requires solving a linear system with only a limited number of degrees of freedom.

We are now in position to substantiate our claim in the introduction, where we briefly mentioned that, in the stochastic setting, a direct application of the MsFEM to approximate the solution to (1.2) isunpractical. It would indeed lead to compute, foreach realizationofAε(x, ω), first a basis set and second a macroscale solution.

This approach has been briefly examined theoretically in [21]. It is prohibitively expensive. We therefore turn to an alternate approach.

2.2. A weakly stochastic setting

We now restrict the general setting and propose a dedicated, practical MsFEM type approach. Following up on previous works (see [5,13,24,41]) and as announced in (1.3), we assume here that the random matrix Aε(x, ω) in (1.2) is aperturbation of a deterministic matrix, in the sense that

Aε(x, ω)≡Aεη(x, ω) =Aε0(x) +η Aε1(x, ω), (2.12) whereη∈Ris a small deterministic parameter,Aε0andAε1are bounded matrices, andAε0is coercive, uniformly inε. We also assume that the matrixAεη itself satisfies the coercivity and boundedness assumptions, uniformly in ηand ε(we refer to [6–8] and [15,25] for other perturbative settings).

The principle of the proposed approach is to compute the MsFEM basis set of functions with the deter- ministic part Aε0 of the matrix Aεη, and next to perform Monte−Carlo realizations for the macroscale prob- lem (1.2)−(2.12), where we keep the exact matrix Aεη (and not only its deterministic part). Following the approach sketched in Section 2.1, we first solve (2.7) with Aε(x) ≡Aε0(x), and build the deterministic finite dimensional space

Wh:= span(φεi, i= 1, . . . , L)

(10)

ET AL.

following (2.8)−(2.9). We next proceed with a standard Galerkin approximation of (1.2)−(2.12) usingWh. For eachm∈ {1, . . . , M}, we consider a realizationAε,mη (·, ω) and computeumS(·, ω)∈ Whsuch that

∀v∈ Wh,

K

K(∇v(x))TAε,mη (x, ω)∇umS(x, ω) dx=

Df(x)v(x) dx. (2.13) Since the MsFEM basis functions are only computed once (rather than for each realization ofAεη(x, ω)), a large computational gain is expected, and obtained, in comparison to the direct approach described above.

3. Numerical simulations

This section is devoted to the many numerical simulations we have performed. We first discuss some imple- mentation details. Next, we numerically estimate the performance of our approach on various test cases, and assess its sensitivity with respect to the magnitude of η. We consider in Section3.2 a test case in dimension one. In Section3.3, we next study two test cases in dimension two. We also study how the presence inAε1(the random component of the matrixAεη) of high frequencies that are not present in the deterministic component Aε0, and that are thus not encoded in the highly oscillatory basis functions, affects the accuracy of our approach.

In Section3.4, we eventually compare our approach with a fully deterministic approach.

Let uεη be the reference solution to (1.2)−(1.3) obtained using a finite element method with a mesh size adapted to the small scale ε,uS be the approximation given by our approach (described in Sect.2.2) anduM be the approximation given by the direct approach (in which the MsFEM basis set is recomputed for each realization Aε,mη (x, ω), as explained at the end of Sect. 2.1). Our goal is to compare the error uS −uεη of our numerical approximation with the erroruM −uεη of the direct and expensive approach. When η is small, we expect the approximationuS to be essentially as accurate as the approximationuM, and we show below that this is indeed the case.

In the sequel, we assess the accuracy using the estimators eL2(u1, u2) =E

u1−u2L2(D)

u2L2(D)

and eH1(u1, u2) =E

u1−u2H1 u2H1 h

h

, (3.1)

whereu1 andu2 are the solutions obtained with any two different methods, and uH1

h:=

K∈Th

u2H1(K)

1/2

(3.2) is the broken H1 norm. The expectation is in turn computed using a Monte−Carlo method. Considering M realizations {Xm(ω)}1≤m≤M of a random variable, e.g. X(ω) = u1(·, ω)−u2(·, ω)H1

u2(·, ω)H1 h h

, we compute the empirical meanμM and the empirical standard deviationσM as

μM(X) = 1 M

M m=1

Xm(ω), σM2 (X) = 1 M 1

M m=1

(Xm(ω)−μM(X))2. (3.3) As a classical consequence of the Central Limit Theorem, the following estimate is commonly employed:

|E(X)−μM(X)| ≤1.96 σM(X)

√M ·

It provides a practical evaluation ofE(X) from the knowledge ofμM(X) andσM(X). The numerical parameters have been determined by an empirical study of convergence. For instance, for the reference solution, we choose

(11)

the mesh sizehsuch that the quantity uε,hη −uε,h/2η H1(D)

uε,h/2η H1(D) is smaller than 0.03%, thereby formally admitting that the approximation has converged inh. The MsFEM parameters are determined likewise.

All the computations have been performed using FreeFem++ [33], with the MPI tools.

3.1. Implementation details

In the deterministic version of the MsFEM, the same matrix Aεappears in the definition (2.7) of the basis functions and in the macroscale variational formulation (2.11). This can be used to expedite the computation of the stiffness matrix associated with (2.11). In our approach, described in Section2.2, the matrix that appears in the definition of the basis functions isAε0, whereas the macroscale variational problem involvesAεη≡Aε0+ηAε1. An additional numerical computation is thus needed.

To solve (2.13), we need to compute, for each element Kand each realizationAε,mη (x, ω), the integrals Kη,mij (ω) =

K

∇φε,Ki (x) T

Aε,mη (x, ω)∇φε,Kj (x) dx, (3.4) where φε,Ki are deterministic functions. We recall that Aεη(x, ω) =Aε0(x) +ηAε1(x, ω) (see (2.12)). To allow for an efficient evaluation of (3.4), we assume henceforth thatAε1is of the form

Aε1(x, ω) =

k∈Zd

1Q+k x

ε

Xk(ω)Bkε(x), (3.5)

where Q= (0,1)d, where (Xk)k∈Zd are scalar random variables, and for anyk Zd, x→Bkε(x) Rd×d are some deterministic functions. We comment on this assumption in Remark3.1below. The important consequence of (3.5) is that we can write the integral (3.4) as a linear combination ofdeterministic integrals over cells of size ε, withrandom coefficients. To simplify the notation, we assume that the spatial dimension isd= 2. We define

p=

min yi

ε,yj ε,yk

ε

, q=

max yi

ε,yj ε,yk

ε

+ 1,

where yi, yj and yk are the y-axis coordinates of the three vertex of K (see Fig. 2). We likewise define the integersl andm(see Fig. 2). We can then write (3.4) as

Kijη,m(ω) =

K

∇φε,Ki (x) T

Aε,mη (x, ω)∇φε,Kj (x) dx=K0,Kij +η

q−1

α=p m−1

β=l

Xα,βm (ω)K1,Kαβij, (3.6) where

K0,Kij =

K

∇φε,Ki (x) T

Aε0(x)∇φε,Kj (x) dx, (3.7)

K1,Kαβij =

(α+1)ε

αε

(β+1)ε

βε

1K(x)

∇φε,Ki (x) T

Bα,βε (x)∇φε,Kj (x) dx. (3.8)

We thus compute once the deterministic integrals (3.7) and (3.8). Next, for each realization ofAεη, we evaluate the stiffness matrix elementsKijη,m(ω) using the right hand side of (3.6). No numerical quadrature is needed. As a consequence of (3.5), most of the work for assembling the stiffness matrix is only performed once, independently of the number of Monte Carlo realizations. This significantly contributes to the gain in term of computational cost.

(12)

ET AL.

Figure 2. To practically compute the integral (3.4), we write that each elementK (here in dimensiond= 2) is a subset of a quadrangle (here [lε, mε]×[pε, qε]) composed of cells of sizeεd.

Remark 3.1. Assumption (3.5) is quite general, and already covers many interesting cases in practice. As explained above, the point in (3.5) is thatAε1 is adirect product(or here, a sum of direct products) of a function depending on xwith a random variable that only depends onω. Otherwise stated, Aε1(x, ω) depends linearly, in an explicit way, ofω. A similar assumption is made when applying reduced basis methods [45] to a problem of the type

Find uλ such that, for anyv,a(uλ, v;λ) =b(v), (3.9) where a(·,·;λ) is a bilinear form parameterized by λ. Assume this problem has been solved for some values i}Ii=1of the parameter, yielding the functions{uλi}Ii=1. Under the assumption thata(·,·;λ) =a0(·,·)+λa1(·,·) (namely, a(·,·;λ) depends linearly on λ), one can precompute the stiffness matrix elements a0(uλi, uλj) and a1(uλi, uλj) for any 1 i, j I. This allows to next perform a very efficient Galerkin approximation of the problem (3.9) (for anyλ) on the space Span(uλi, i= 1, . . . , I).

3.2. One-dimensional test-case

The purpose of this section is threefold. We first calibrate the numberM of realizations considered for the Monte−Carlo method for the two-dimensional numerical experiments that we consider in the sequel. We next investigate how the accuracy of our approach depends onη and on the presence of frequencies in the random coefficientaεη that are not taken into account in the MsFEM basis set functions. The low computational costs that we face in this one-dimensional situation allow us to test our approach more comprehensively than in the two-dimensional test-cases described below.

Let (Xk)k∈Z denote a sequence of independent, identically distributed scalar random variables uniformly distributed in [0,1]. We consider the random coefficient

aεη(x, ω) =

k∈Z

1(k,k+1]

x

ε 5 + 50 sin2 πx

ε

+ηXk(ω)κ sin2 ζπx

ε

,

(13)

0 5 10 15 20 25 30 35 40 45 50

−2 0 2 4 6 8 10 12 14 16x 10−4

M

Figure 3. Convergence of the indicatoreH1(uM, uεη) (see (3.1)), forη= 1,ζ= 1 andκ= 55.

For each value ofM, we plot the empirical mean along with its confidence interval, computed from the firstM realizations. We only plot the results for the first 50 realizations.

which is a particular example of the expansion (2.12) with aε0(x) = 5 + 50 sin2

πx ε

and aε1(x, ω) =

k∈Z

1(k,k+1]

x ε

Xk(ω)κsin2 ζπx

ε

, and that satisfies the structural assumption (3.5). We setε= 0.025 and chooseκsuch that the quantity

R(κ, ζ) = aε1

aε0

L(D×Ω)

= SupEssω∈Ω aε1(·, ω)

aε0

L(D)

(3.10) has the same valueR(κ, ζ) = 1 for the three different values ofζ={1,3,7}we consider below. This yields the choices (κ, ζ) = (55,1), (κ, ζ) = (14.38,3) and (κ, ζ) = (8.39,7). We analytically compute the reference function uεη, solution to

d dx

aεη(x, ω)duεη dx(x, ω)

= 1 in (0,1), uεη(0, ω) =uεη(1, ω) = 0,

as well as the MsFEM basis functions for both approaches. LetuM anduS be the approximation ofuεη by the two MsFEM approaches described above, where the coarse mesh size ish= 1/30.

We first calibrate the number of independent realizations to accurately approximate the exact expectation in (3.1) by the empirical mean (3.3). To this aim, we present on Figure3 the mean and the confidence interval computed using (3.3) for an increasing numberM of realizations (we compute up to 1000 independent realiza- tions). We check that this indicator reaches a plateau forM 30, and thus converges fast. On this example, considering 30 realizations is hence sufficient to accurately compute the error (3.1). Based on this observation, we will only considerM = 30 realizations in the two dimensional examples of Section3.3.

Remark 3.2. There is no reason to think that the calibration of our parameters that we perform in the one- dimensional situation provides an adequate adaptation of these parameters for the higher dimensional setting.

Références

Documents relatifs

Problems dealing with wave scattering in a heterogeneous medium are investigated in applications, whose associated scales can be extremely different (seismic waves, ultrasound

Applications of the mortar element method are also present in situations where domain décomposition is not (only) involved for parallel purpose but has been used to mesh the

We consider the Finite Volume method for conservation laws with high order polynomial reconstruction in the context of one dimensional problems.. A particular reconstruction based on

Also, by using some precise behaviour and decay estimates of the sequence of the eigenvalues of the Sinc kernel operator, we give fairly good estimates of the number of degrees

It follows that stability of a steady state equilibrium implies that, locally, there will exist initial values for government debt for which a policy of rolling over the debt

In addition, we also derive several new results: optimal error estimates are derived for Q ℓ -rectangular FEs (for such elements we don’t know how to obtain error estimates using

Our aim in this paper is to study the convergence of approximate solutions to the resulting coupled problem obtained with both a linear finite element method and a cell centred

We show that the orthogonal free quantum groups are not inner amenable and we construct an explicit proper cocycle weakly contained in the regular representation.. This strengthens